Vous êtes sur la page 1sur 8

J. Am. Ceram. Soc.

, 84 [6] 126572 (2001)

journal

Effect of Substitution of Manganese for Iron on the Structure and


Electrical Properties of Yttrium Ferrite
Xueqiang Cao
College of Chemistry and Chemical Engineering, Hunan Normal University, Changsha, Hunan 410081, China

Chan-Soo Kim and Han-Ill Yoo*


Department of Inorganic Materials Engineering, Seoul National University, Seoul 151742, Korea
The solubility of manganese in YFe1yMnyO3 reaches y 0.4.
Among these compositions, YFe0.6Mn0.4O3 has the highest
conductivity with 2.16 1cm1 (1000C, air), which is
>1 order of magnitude higher than that of yttrium ferrite,
YFeO3 ( 0.10 1cm1, 1000C, air). In the range 0.2 <
y < 0.4, the conductivity is linearly proportional to the
manganese content. The mechanism of direct transport of
polarons among manganese sites (y > 0.2) is proposed to
explain the electrical behavior of manganese-doped samples.
X-ray diffractometry and infrared spectroscopy have been
applied for the structural characterization of the samples.
I.

In this work, we attempted to improve the conductivity of


YFeO3 by substituting calcium for yttrium and manganese for iron.
However, the result was opposite to the expectation. The substitution of calcium for yttrium in YFe1yMnyO3 resulted only in a
small increase of conductivity that was much lower than that of
Y0.9Ca0.1FeO3. This interesting phenomenon has attracted our
attention to study the difference between the conduction mechanism of pure YFeO3 and that of Ca,Mn-doped YFeO3.
II.

Experimental Procedure

The main chemicals used in this work were analytical-reagentgrade (99.9%). YFe1yMnyO3 specimens were prepared using
the Pechini method12 or solid-state method. A stoichiometric
mixture of Y(NO3)3, Fe(NO3)3, and MnCO3 was combined with
citric acid and ethylene glycol. The polymer solution was evaporated, dried, burned at 500C, and calcined at 900C for 10 h. The
product was milled using zirconia balls for 24 h and then was cold
isostatically pressed and sintered at 1350C for 15 h. If the
solid-state method was applied, mixtures of corresponding metal
oxides or carbonates in the proper ratios were directly used for
calcination and sintering; the remainder of the preparation procedure was similar to the above. Samples that were prepared using
the Pechini method usually had a slightly higher electrical conductivity (20%) than those prepared using the solid-state method
because of the high homogeneity and good contact of the particles.
Phase impurity and crystal structure of each composition were
examined using powder X-ray diffractometry (XRD; Model XRD2000, Scintag, Inc., Santa Clara, CA) using CuK1 radiation.
Lattice parameters were calibrated using an internal standard of
silicon single crystal with a scanning rate of 1/min. Fouriertransform infrared spectra (FT-IR) were recorded using KBr
pellets and a spectrophotometer (Model 1725X8700, Perkin
Elmer Co., Norwalk, CT) from 10 000 to 400 cm1. Direct-current
electrical conductivity measurement was performed using a fourprobe technique programmable current source (Model 200,
Keithley Instruments, Inc., Cleveland, OH) and a multimeter
(Model 2000, Keithley Instruments). The Seebeck coefficient was
determined using a temperature gradient of maximum 10 K along
the length of the specimen and repeated measurements of the
temperature difference and the corresponding Seebeck voltage. At
least 20 points were measured, and the slope of the resulting line
was the Seebeck coefficient. The PO2 was controlled using a gas
mixture of N2/O2 or CO2/CO and monitored using a YSZ oxygen
sensor. The specimen size was 1.8 mm 1.8 mm 20 mm.

Introduction

HE perovskite-structured lanthanum transition-metal oxides


doped with alkaline-earth metals, i.e., La1xAxMO3 (A Ca,
Sr and M Mn, Fe, Cr, Co) are of technological importance for
their use in solid-oxide fuel cells (SOFCs), chemical sensors, and
catalysts.1,2 One of the most promising cathode materials for
SOFC is La1xSrxMnO3, even though it has some disadvantages,
namely formation of high-resistance La2Zr2O7 and SrZrO3 with
Y2O3-stabilized ZrO2 (YSZ), chemical instability at high temperature under low oxygen partial pressure (PO2), and diffusion of
manganese into YSZ.3 Y0.9Ca0.1FeO3 is a newly proposed cathode
material for SOFCs and has been extensively studied.1,4 Its
conductivity is much higher than that of YFeO3 but not high
enough for practical application. Y1xCaxMnO3 has high electrical
conductivity; however, a microcracking problem and low stability
make it useless in SOFCs.3,5 The search for materials possessing
the structural stability and electrical conductivity required in a fuel
cell environment has led to the investigation of substitutionally
mixed systems, whose end members possess material properties
desired in the substituted materials. Recently studied examples are
Y(Cr,Mn)O3,6 La(Cr,Mn)O3,6,7 (La,Sr)(Cr,Mn)O3,8 (La,Sr)(Co,Fe)O3,9 and (La,Sr/Ca/Ba)(Co,Fe)O3.10,11 The mixed systems, however, can have properties much different from those of
either of the end members. For example, the substitutional inclusion of a small amount of the ion associated with the more
conductive end member leads to a conductivity that decreases
orders of magnitude below that which occurs for either end
member of the series.6 8

H. U. Andersoncontributing editor

III.
Manuscript No. 189224. Received July 15, 1999; approved March 13, 2000.
Supported by the New and Renewable Energy Development Center, Korea Energy
Management Corporation.
*Member, American Ceramic Society.

To whom all correspondence should be addressed. Temporary address: Forschungszentrum Julich, D-52425 Julich, Germany.

Results and Discussion

(1) Crystal Structure of Y1xCaxFe1yMnyO3


The lattice parameters of Y1xCaxFe1yMnyO3 are summarized
in Table I with the ionic radii of Y3, Ca2, Fe3, and Mn3.
The lattice parameter change does not follow Vegards law.
1265

1266

Journal of the American Ceramic SocietyCao et al.

Table I. Lattice Parameters of Y1xCaxFe1yMnyO3


x

0
0
0
0
0
0
0
0.1
LaFeO3
LaMnO3

0
0.1
0.2
0.3
0.35
0.4
1
0.4

a ()
( 0.003)

b ()
( 0.006)

c ()
( 0.005)

V (3)
( 0.4)

D (%)

5.282
5.279
5.280
5.276
5.270
5.271
5.24
5.271
5.567
5.537

5.598
5.599
5.604
5.612
5.643
5.651
5.84
5.569
5.553
5.741

7.602
7.589
7.585
7.548
7.534
7.517
7.36
7.537
7.855
7.694

224.8
224.3
224.4
223.5
224.0
223.9
225.2
221.2
242.8
244.6

11.51
11.51
11.52
11.50
11.64
11.65
12.33
11.33
9.83
8.00

Ionic radii (Shannon15): O2 1.40 , Ca2(CN12) 1.34 , Y3(CN12)


1.18 , Fe3(CN6) 0.64 , and Mn3(CN6) 0.66 . a (abc/21/2)1/3, a1
3
a, a2 b, a3 c/21/2, D (i1
ai a /ai)/3 100. Reference 17. Reference 18.

Powder XRD analysis confirms that the compounds with x 0.1


and y 0.4 are single phase with the orthorhombic structure of
YFeO3 (see Fig. 1). When y 0.8, the product keeps the
hexagonal structure of YMnO3. Products with 0.4 y 0.8 are
mixtures of the above two phases. Y0.9Ca0.1Fe0.6Mn0.4O3 is also

Fig. 1.

XRD patterns of YFe1yMnyO3.

Vol. 84, No. 6

single phase. It is reasonable to suppose that Ca2 and Mn3 are


introduced to the yttrium and iron sites, respectively, because of
their similar radii and similar electronic states. YFeO3 is orthorhombic with space group Pbnm.13,14 Y3 is surrounded by 12 O2,
and Fe3 is surrounded by 6 O2, forming an FeO6 octahedron.
YMnO3 is hexagonal with space group P63cm.15 In this structure,
Y3 has a sevenfold coordination with O2 and Mn3 fivefold
coordination forming a MnO5 bipyramid. The structural mismatch
prevents YFeO3 and YMnO3 from forming a complete solid
solution in the entire compositional range.
The lattice parameter change trend is shown in Table I. The
cubic perovskite (formula ABO3) adjusts to t 1 (Goldschmidt
tolerance, t (rA rO)/21/2(rB rO)) by a cooperative rotation
of a BO6 octahedron that buckles the BOB bond angle from
180 to (180 ), where is the tilting angle of BO6 in the (100)
plane, and it increases with decreasing t to lower the space group
symmetry from cubic to tetragonal, rhombohedral, or orthorhombic. The t-factor of YFeO3 is 0.89, and it requires an orthorhombic
perovskite structure by the rotation of an FeO6 octahedron around
the orthorhombic b-axis. With the increase of manganese content
in YFe1yMnyO3, the BO bond distance increases because of the
larger radius of Mn3 compared with Fe3. Longer BO distance
decreases the t-factor, and the BO6 octahedron tilts more, resulting
in a decrease of lattice parameters a and c, and the increase of b.
The distortion D increases with the substitution of manganese for
iron. The change in lattice parameters from YFeO3 to
YFe0.6Mn0.4O3 is quite similar to that from LaFeO3 to LaMnO3,
implying that Mn3 exists in the B-site of YFeO3. Ca2 is larger
than Y3. The replacement of Y3 by Ca2 results in the increase
of t-factor and decrease of distortion.
Figure 2 shows the IR spectra of YFe1yMnyO3. The regular

Fig. 2.

FT-IR spectra of YFe1yMnyO3.

June 2001

Effect of Substitution of Mn for Fe on the Structure and Electrical Properties of Yttrium Ferrite

cubic perovskite structure belongs to the O1h space group. According to group theory, the optically active lattice vibrations are given
by 3F1u F2u,16 of which the F2u mode is inactive in the IR
region. Therefore, the IR spectra of the perovskite structure should
show three absorption bands. These can be attributed to the BO
bond stretching vibration of BO6 octahedra (1:F1u), the bending
vibration of the BO bond (2:F1u), and the lattice vibration
(3:F1u). The expected energy order is usually 1 2 3. Any
deviation from the cubic symmetry causes splitting of these three
bands. Under our experimental condition, we can observe only the
former two bands. Figure 2 shows that the substitution of manganese for iron makes the absorption peak shift toward the higherenergy side, implying that the covalence of the BO bond is
increased and the ionicity is decreased. On the other hand, the
substitution of manganese for iron also splits band 2 in two peaks.
Especially for YFe0.6Mn0.4O3, two absorption peaks in the 2 band
can be distinguished clearly, indicating that the structure deviates
more and more from the cubic symmetry, as proved by XRD
results.
(2) Electrical Properties
(A) YFeO3: Figure 3 shows the electrical conductivity ()
and Seebeck coefficient () of YFeO3 measured in an air atmosphere. In the entire experimental temperature range, YFeO3 has the
n-type conduction character. The Seebeck coefficient and electrical conductivity can be calculated using the following equations:1,8

k
21 C
ln
S0
e
C

(1)

a 2 e 2 N mC1 C
expEa/kT
kT

(2)

where e is the absolute value of the electronic charge, a the


intersite distance for small polaron hopping (3.8 1010 m), Nm
the density of possible sites (1.8 1028 m3), C the charge-carrier
concentration, (1 C) the available sites for the jumping of charge
carriers, the vibrational frequency (or hopping frequency) of

Fig. 3.

1267

polarons, k the Boltzman constant, T the absolute temperature, Ea


the activation energy, and S0 the vibration entropy per particle
(often neglected because of its small value).1 The factor 2 takes
into account the spin degeneracy. KrogerVink notation is used in

this paper: FeFe


means Fe3, FeFe means Fe2, and FeFe
means
4
3
3
Fe ; when Fe sites in YFeO3 are replaced by Mn , Mn2,

and Mn4; then MnFe


means Mn3, MnFe means Mn2, and MnFe
4
means Mn For YFeO3, Eqs. (1) and (2) should be expressed as

2FeFe
k
21 2c
k
ln
ln

e
e
c
FeFe

a 2 e 2 N mc1 2c
expEa/kT
kT

(3)

(4)

Because YFeO3 has n-type conduction, it is reasonable to


suppose that electrons associated with the formation of Fe2 or

FeFeFeFe
are charge carriers. In Eqs. (3) and (4), c is the
concentration of FeFe and (1 2c) is the available sites for the
jumping of charge carriers, and it is equal to the concentration of
FeFe, which can be easily obtained from the disproportionation

equation, 2FeFe
FeFe FeFe
.
As calculated using the above equations, at 1000C, c 0.03
and (5.7 5.6) 1014 Hz, which is slightly dependent on the
temperature.
The electrical conductivity of YFeO3 versus PO2 is shown in
Fig. 4. At 1000C, YFeO3 is stable until log PO2 13.5. The
conductivity behavior of YFeO3 in this study is somewhat different from the former work,1,4 but the Ca-doped YFeO3 has the same
electrical property as the former. In this study, YFeO3 does not
show a conductivity minimum. This difference is possibly induced
by different specimen preparation conditions, such as contamination and nonstoichiometric composition. Because the undoped
YFeO3 of this work shows the n-type conduction, it is possible that
the formation of FeFe is electronically compensated by the existence of impurities with valence 4, such as Zr4 and Si4, which
might be obtained during ball-milling. This possibility has been
checked using inductively coupled plasma spectrometry (ICP).

Electrical conductivity and Seebeck coefficient of YFeO3 in an air atmosphere.

1268

Fig. 4.

Journal of the American Ceramic SocietyCao et al.

Electrical conductivity of YFeO3 as a function of PO2.

However, the result has proved that impurity levels of Zr4 and
Si4 were below the sensitivity limit of the ICP instrument
( 0.01 wt%, approximately equal to 0.02 mol% for Zr4 and
0.07 mol% for Si4). These values are dramatically lower than the
charge-carrier concentration, 3 mol%. Therefore, the possibility of
contamination to make the undoped YFeO3 exhibit n-type conduction can be eliminated. On the other hand, a minor deviation in the
stoichiometry, if it exists, should not affect the conduction type
because Y3 and Fe3 have the same valence, and no electronic
compensation occurs. It has been observed that a trace of garnet
Y3Fe5O12 exists in the system Y1xCaxFeO3 (x 0 0.1), but
YFeO3 exhibits p-type conduction,1 implying that stoichiometry
deviation does not determine the conduction type. Therefore, it is
more reasonable to suppose the n-type conduction of YFeO3 to be
intrinsic than to be extrinsic.
(B) YFe1yMnyO3: Figure 5 shows the electrical conductivity of YFe1yMnyO3 versus temperature in an air atmosphere.
These curves are best fitted with s 1 in the equation
(C/T s) exp(-Ea/kT) for other than s 0 or s 3/2, implying that
electrical conduction occurs through the adiabatic small-polaron
hopping mechanism. The dotted lines in Fig. 5 are a theoretical
fitting to the experimental data according to this equation. The
conductivity of YFe1yMnyO3 at selected temperatures from 700
to 1200C is shown in Fig. 6. At 1000C, the conductivity of
YFe0.6Mn0.4O3 is 2.16 1cm1, 1 order of magnitude higher
than that of YFeO3 (0.10 1cm1). The activation energy of
YFe1yMnyO3 is shown in Fig. 7.
Figure 6 shows that the electrical conductivity of YFe1yMnyO3
can be clearly distinguished into two regimes with y 0.2 as the
crossover point. In the range y 0.2, the conductivity decreases
first and then increases with y 0.05 as the minimum point. When
y 0.2, the conductivity increases in linear proportion to the
manganese content, and, in this range, the activation energy is
almost constant. The conductivity behavior of YFe1yMnyO3 is
similar to that of La(Cr,Mn)O3, whose conductivity minimum also
occurs at Mn 0.05, and can be well described with the
random-well and random-barrier models.7,8 Despite the comparatively higher intrinsic conductivity of YMnO3, a small substitution
of manganese for iron in YFeO3 results in a decrease of conductivity below 900C. This decrease becomes larger as temperature
decreases. In the system YFe1yMnO3, the manganese site is

Vol. 84, No. 6

Fig. 5. Electrical conductivity of YFe1yMnyO3 in an air atmosphere


((- - -) theoretical fitting to the data).

Fig. 6. Electrical conductivity of YFe1yMnyO3 in an air atmosphere


((- - -) where open symbols are theoretical values).

supposed to be energetically lower than the iron site. The activation energy of YFe0.95Mn0.05O3 (Ea 1.02 eV) must represent the
energy for a small polaron at a manganese site to hop to a
neighbored iron site (YFeO3, Ea 0.82 eV). The energy difference between the manganese site and the iron site makes the
manganese sites (y 0.05) act as traps for polarons migrating
among the iron sites. The trapping mechanism can be expressed

as FeFe MnFe
3 FeFe
MnFe. Small polarons (FeFe
FeFe
)
are supposed to be the main charge carrier in the systems

June 2001

Effect of Substitution of Mn for Fe on the Structure and Electrical Properties of Yttrium Ferrite

1269

particles trapped in an extended cluster of manganese traps no


longer need to make a transition to an iron site to participate in the
conduction; in other words, direct small polaron transport among
manganese sites occurs. The manganese content limit of this
crossover phenomenon is y 0.2 and Mn 0.3 in the system
La(Cr,Mn)O3. 0.2 y 0.05 is the crossover region. Within this
region, the conductivity increases slowly with manganese content,
and the activation energy decreases.
Figure 8 shows the thermopower and electrical conductivity of
YFe0.6Mn0.4O3 measured in an air atmosphere. Because of the low
energy difference between Mn4 and Mn3, the electrical conduction in the Mn3-containing material occurs normally between
Mn3 and Mn4, which has been proved in many works.20 The
charge carriers are reasonably supposed to be holes associated with
Mn4 below 950C (p-type) and electrons associated with Mn3
above 950C (n-type).
The charge-carrier concentration and vibrarional frequency of
polarons can be calculated using the following equations:

2Mn3
c
0.42c

Mn2

0.20.5c
c

Mn4
0.20.5c
c

T 950C
T 950C
(5)

2Mn4
k
20.20.5c
k
ln
ln
e
Mn3
e
c

T 950C
Fig. 7.

Activation energy of YFe1yMnyO3.

Y1xCaxFeO31 and La1xSrxFeO319; for YFeO3 of this work,

electrons (FeFeFeFe
), as discussed above, are charge carriers, and,

therefore, the oxidation of FeFe to FeFe


reasonably decreases the
electrical conductivity. As the manganese concentration increases,
a connected path of manganese sites is formed. Once this occurs,

Fig. 8.

k
e

ln

2Mn3
k
20.42c
ln
Mn4
e
c

a 2e 2N m
c0.20.5c expEa/kT
kT

a 2e 2N m
c0.42c expEa/kT
kT

Electrical conductivity and Seebeck coefficient of YFe0.6Mn0.4O3 in an air atmosphere.

(6)

T 950C
(7)
T 950C
T 950C

(8)

(9)

1270

Journal of the American Ceramic SocietyCao et al.

Fig. 9. ln KD as a function of temperature in an air atmosphere (solid


symbols for YFe0.6Mn0.4O3 and open symbols for Y0.9Ca0.1Fe0.6Mn0.4O3).

Equations (5)(9) are not suitable for the change regime of


conduction type (950C). Within this regime, electrons and holes
contribute to the thermopower and electrical conduction. In the
following calculation of Gibbs free energy, enthalpy, and entropy
of Mn3 disproportionation from the plot of ln KD versus inverse
temperature, one or two points around the type-change point are
neglected.
In YFe0.6Mn0.4O3, [Mn4] [Mn2] and oxygen vacancies in
the air atmosphere are not taken into account. The doping of
manganese for iron does not change the crystal structure of YFeO3
on an obvious level, as proved by XRD measurement (see Table I).
Therefore, a and Nm have the same values as in YFeO3. The
calculation results for 1000C are c 0.19 and 5.62 10.14
The value (5.89 0.04) 1014 exceeds the expected optical
phonon frequency (1013 Hz), and it is approximately equal to
that of Mn3O4 ((5.1 0.4) 1014).20 Similarly high values
have been reported for Fe3O421 and for cobalt and manganese
ferrites.22,23
The Gibss free energy (GD), enthalpy (HD), and entropy (SD) of
the disproportionation Eq. (5) have the following relation:

HD 1 SD
GD

ln KD
RT
R T
R

(10)

where KD is the disproportionation constant. The plot of ln KD vs


1/T should result in a straight line, as shown in Fig. 9. There is an
abrupt change in this plot because of the change of conduction
type. (HD) and (SD) for the p-type and n-type conduction ranges
are listed in Table II. For Mn3O4 (p-type), (HD) 26.21 kJmol1

Table II. Disproportionation Enthalpy and Entropy of


2Mn3 Mn2 Mn4
Composition

YFe0.6Mn0.4O3
YFe0.6Mn0.4O3
Y0.9Ca0.1Fe0.6Mn0.4O3
Y0.9Ca0.1Fe0.6Mn0.4O3

T (C)

950
950
920
920

HD (kJmol

66 4
97.1 0.9
24 1
19.6 0.2

SD (Jmol

42.3 0.1
91.8 0.6
4.62 0.03
22.5 0.1

Vol. 84, No. 6

Fig. 10. Electrical conductivity of YFe0.6Mn0.4O3 as a function of PO2


(() is decomposition point).

) 17.46 Jmol1K1.22
and (SD
Because the electrical conduction occurs only among the
manganese sites, and the function of iron sites can be neglected
when 0.4 y 0.2, we have two reasonable assumptions in
the conductivity calculation of compositions other than
YFe0.6Mn0.4O3: (i) (HD) and (SD) are independent of composition;
and (ii) compositions with 0.4 y 0.2 have similar activation
energy and similar vibrational frequency.
The theoretical fitting of conductivity is shown in Fig. 6. There
is a good agreement between the theoretical values and experimental results, further proving that the conduction occurs directly
among manganese sites when y 0.2.
The electrical conductivity of YFe0.6Mn0.4O3 as a function of
PO2 at 900, 1000, and 1100C is shown in Fig. 10. For all these
isotherms, conductivities are almost constant within a certain range
of PO2 depending on the temperature. These isotherms do not
exhibit the oxygen-activity-dependence behavior. The same phenomenon has been observed in the system Y1xCaxMnO3 by
Stevensen et al.5 The low phase stability of Y-Mn-O oxides can be
attributed to the small size of Y3, which apparently decreases the
stability of yttrium-containing perovskite structures, so that they
dissociate as soon as the loss of oxygen begins. Lanthanum
compounds with perovskite structures, such as LaMnO324 and
LaFeO3,25 have much higher stability than that of yttriumcontaining compounds, and the dependence of conductivity on PO2
can be observed clearly.5,19,26 When YFe0.6Mn0.4O3 begins to
dissociate, its conductivity increases sharply. As proved by XRD
experiment and indicated by Stevensen et al.,5 Mn3O4 is the stable
manganese oxide under the PO2 range where Y1xCaxMnO3 begins
Table III. Gibbs Free Energies for the Formation of
YFe0.6Mn0.4O3 and Y0.9Ca0.1Fe0.6Mn0.4O3
Gibbs free energy, Gf (kJmol1)
900C

1000C

1100C

Composition

20.2 0.7 13.5 0.5 10.5 0.8 YFe0.6Mn0.4O3


18 1
16.5 0.7 6.2 0.7 Y0.9Ca0.1Fe0.6Mn0.4O3

June 2001

Fig. 11.
phere.

Effect of Substitution of Mn for Fe on the Structure and Electrical Properties of Yttrium Ferrite

Electrical conductivity of Y1xCaxFe1yMnyO3 in an air atmos-

to dissociate (log PO2 8 to 9 at 1000C). This pressure range


is similar to that of YFe0.6Mn0.4O3 (log PO2 7 to 9). The
Gibbs free energies for the formation of YFe0.6Mn0.4O3 are listed
in Table III. These values are much higher than those of (La,Ca/
Sr)MnO3,5 indicating a low phase stability of (Y,Ca)(Fe,Mn)O3.

Fig. 12.

1271

(C) Y0.9Ca0.1Fe0.6Mn0.4O3: The conductivity of calciumdoped samples is shown in Fig. 11. The conductivity increases
with the addition of calcium, as expected. Y0.8Ca0.2Fe0.6Mn0.4O3
has almost the same conductivity as Y0.9Ca0.1Fe0.6Mn0.4O3. The
former is not a single phase, and it should have similar chargecarrier concentration to Y0.9Ca0.1Fe0.6Mn0.4O3, because Ca 0.1
is the doping limit, as proved by XRD experiment. However, the
most curious phenomenon is that the conductivity of
Y0.9Ca0.1Fe0.6Mn0.4O3 (5.6 1cm1, 1000C) is much lower
than that of Y0.9Ca0.1FeO3 (18.0 1cm1, 1000C). It is
supposed that the electronic compensation of calcium doping in
Y0.9Ca0.1FeO3 is the formation of Fe4 and that the sample has
p-type conduction. In Y0.9Ca0.1Fe0.6Mn0.4O3, the electronic compensation must be the formation of Mn4. In CaMn1xFexO3y
and SrMn1xFexO3, all iron sites are in valence Fe3, and most
of the manganese sites are in Mn4.27,28
The thermopower of Y0.9Ca0.1Fe0.6Mn0.4O3 in an air atmosphere is shown in Fig. 12. Below 920C, it has p-type conduction,
and, above 920C, it changes to n-type. Its Seebeck coefficient is
in a good linear proportion with inverse temperature, but there is
a curvature at the change of the conduction mechanism, the same
as YFe0.6Mn0.4O3. On the other hand, this change coincides with
the curvature of the electrical conductivity plot. The n-type
conduction has a slightly higher activation energy (0.422 0.004
eV) than that of p-type conduction (0.326 0.003 eV). The
conductivity calculation of Y0.9Ca0.1Fe0.6Mn0.4O3 is similar to
that of YFe1-yMnyO3, and c 0.24, 1.3 1014 (1000C). The
equilibrium constant of Mn3 disproportionation in the calciumdoped sample is shown in Fig. 9. (HD) and (SD) obtained from the
plot of ln KD vs 1/T are listed in Table II.
The electrical conductivity of Y0.9Ca0.1Fe0.6Mn0.4O3 as a function of PO2 is shown in Fig. 13, which also indicates its lower phase
stability than that of YFe0.6Mn0.4O3 because of the addition of
Ca2. These isotherms are very similar to those of YFe0.6Mn0.4O3,
and the Gibbs free energy of formation is also listed in Table III.
The reason for conduction-type change is not clear at this time.
One possible reason is the change of crystal structure at high

Electrical conductivity and Seebeck coefficient of Y0.9Ca0.1Fe0.6Mn0.4O3 in an air atmosphere.

1272

Journal of the American Ceramic SocietyCao et al.

Fig. 13. Electrical conductivity of Y0.9Ca0.1Fe0.6Mn0.4O3 as a function


of PO2 (() decomposition point).

temperature. When the structure of Mn3O4 changes at high


temperature, the conductivity and valence distribution change
abruptly.20
IV.

Conclusion

The substitution of manganese for iron in YFeO3 enhances the


conductivity if Mn 0.2, and it reduces the conductivity when
manganese content is very low (5 mol%). The conduction
mechanism of manganese-doped samples is different from that of
YFeO3 and depends on the doping level. Pure YFeO3 without
doping shows n-type conduction and the charge carrier is assumed
to be electrons associated with Fe2. In the range Mn 0.2,
manganese sites act as an energy trap; with Mn 0.2, a direct
transportation pathway among manganese sites occurs, and this
assumption has been proved by a good fitting of the experimental
data of conductivity.
Acknowledgments
We wish to thank our colleagues Mr. J. Oh. Hong and Mr. C. R. Song of Seoul
National University and Dr. Tietz of Forschungszentrum Julich, Germany, for helpful
discussion and Mr. P. Lersch of Forschungszentrum Julich for part of the XRD
measurements.

References
1
H.-I. Yoo and C.-S. Kim, Electrical Properties and Defect Structure of Y1xCaxFeO3 Othorferrites, Solid State Ionics, 53, 58391 (1992).

Vol. 84, No. 6

2
H.-B. Park, H.-J. Kweon, Y.-S. Hong, S.-J. Kim, and K. Kim, Preparation of
La1xSrxMnO3 Powders by Combustion of Poly(ethylene glycol)Metal Nitrate Gel
Precursors, J. Mater. Sci., 32, 57 65 (1997).
3
B. Fu, W. Huebner, M. F. Trubelja, and V. S. Stubican, (Y1xCax)FeO3: A
Potential Cathode Material for Solid Oxide Fuel Cells; pp. 276 78 in The 3rd
International Symposium in SOFC (Honolulu, HI, May 18 21, 1993). Edited by S. C.
Singhal and H. Iwahara. Electrochemical Society, Princeton, NJ, 1993.
4
C.-S. Kim and H.-I. Yoo, A New Candidate for the Solid Oxide Fuel Cell
Cathode, Y1xCaxFeO3, J. Electrochem. Soc., 143 [9] 286370 (1996).
5
J. W. Stevensen, M. M. Nasrallah, H. U. Anderson, and D. M. Sparlin, Defect
Structure of Y1xCaxMnO3 and La1xCaxMnO3; I. Electrical Properties; II. OxidationReduction Behavior, J. Solid State Chem., 102, 17597 (1993).
6
W. J. Weber, C. W. Griffin, and J. L. Bates, Effects of Cation Substitution on
Electrical and Thermal Transport Properties of YCrO3 and LaCrO3, J. Am. Ceram.
Soc.,70 [4] 26570 (1987).
7
R. Raffaelle, H. U. Anderson, D. M. Sparlin, and P. E. Parris, Evidence for a
Crossover from Multiple Trapping to Percolation in the High-Temperature Electrical
Conductivity of Mn-Doped LaCrO3, Phys. Rev. Lett., 65 [11] 1383 86 (1990).
8
R. Raffaelle, H. U. Anderson, D. M. Sparlin, and P. E. Parris, Transport
Anomalies in the High-Temperature Hopping Conductivity and Thermopower of
Sr-Doped La(Cr,Mn)O3, Phys. Rev. B: Condens. Matter, 43 [10] 799199 (1991).
9
L.-W. Tai, M. M. Nasrallah, H. U. Anderson, D. M. Sparlin, and S. R. Sehlin,
Structure and Electrical Properties of La1xSrxCo1yFeyO3, Part I. The System
La0.8Sr0.2Co1yFeyO3; Part II. The System La1xSrxCo0.2Fe0.8O3, Solid State Ionics,
76, 259 71; 273 83 (1995).
10
J. W. Stevenson, T. R. Armstrong, R. D. Carneim, L. R. Pederson, and W. J.
Weber, Electrochemical Properties of Mixed Conductivity Perovskites La1xMxCo1yFeyO3 (M Sr, Ba, Ca), J. Electrochem. Soc., 143 [9] 272229 (1996).
11
M. H. R. Lankhorst and J. E. ten Elshof, Thermodynamic Quantities and Defect
Structure of La0.6Sr0.4Co1yFeyO3 (y 0 0.6) from High-Temperature Coulometric Titration Experiments, J. Solid State Chem., 130, 30210 (1997).
12
M. P. Pechini, Methods of Preparing Lead and Alkaline-Earth Titanates and
Niobates Using the Same to Form a Capacitor, U.S. Pat. No. 3 330 697, July 11,
1967.
13
P. Coppons and M. Eibschutz, Determination of the Crystal Structure of Yttrium
Othoferrite and Refinement of Gadolinium Othoferrite, Acta Crystallogr., 19,
524 31 (1965).
14
M. Marezio, J. P. Remeika, and P. D. Derrier, The Crystal Chemistry of the
Rare-Earth Othoferrites, Acta Crystallogr. B: Cryst. Chem., 26, 2008 22 (1970).
15
E. Pollert, S. Krupicka, and E. Kuzmicova, Structural Study of Pr1xCaxMnO3
and Y1xCaxMnO3 Perovskites, J. Phys. Chem. Solids, 43 [12] 1137 45 (1982).
16
Y. Wu, Z. Yu, and S. Liu, Preparation, Crystal Structure, and Vibrational
Spectra of Perovskite-Type Oxides LaMyM1yO3 (M, M Mn, Fe, Co), J. Solid
State Chem., 112, 157 60 (1994).
17
H. W. Brinks, H. Fjellvag, and A. Kjekshus, Synthesis of Metastable PerovskiteType YMnO3 and HoMnO3, J. Solid State Chem., 129, 334 40 (1997).
18
Powder Diffraction Files No.371493 (LaFeO3) and No.351353 (LaMnO3).
International Centre for Diffraction Data, Newtown Square, PA.
19
J. Mizusaki, T. Sasamoto, W. R. Cannon, and H. K. Bowen, Electronic
Conductivity, Seebeck Coefficient, and Defect Structure of La1xSrxFeO3 (x 0.1,
0.25), J. Am. Ceram. Soc., 66 [4] 24752 (1983).
20
S. E. Dorris and T. O. Mason, Electrical Properties and Cation Valences in
Mn3O4, J. Am. Ceram. Soc.,71 [5] 379 85 (1988).
21
T. O. Mason and H. K. Bowen, Electronic Conduction and Thermopower of
Magnetite and Iron Aluminate Spinels, J. Am. Ceram. Soc., 64 [4] 237 42 (1981).
22
S. Erickson and T. O. Mason, Nonstoichiometry, Cation Distribution, and
Electrical Properties in Fe3O4CoFe2O4 at High Temperature, J. Solid State Chem.,
59, 4253 (1985).
23
C. Carter and T. O. Mason, Electrical Properties and Site Distribution of Cations
in (MnyCo1y)0.4Fe2.6O4, J. Am. Ceram. Soc., 71 [4] 21318 (1989).
24
T. Nakamura, Isothermal Decomposition of Ternary Oxides AxByOz on an
Isobar-Stability of Perovskite ABO3 (A La, Sm, Dy; B Mn, Fe) in a Reducing
Atmosphere, J. Solid State Chem., 38, 229 38 (1981).
25
N. Kimizuka, A. Yamamot, and H. Ohashi, The Stability of the Phases in the
Ln2O3FeOFe2O3 Systems which are Stable at Elevated Temperatures (Ln: Lanthanide Elements and Y), J. Solid State Chem., 49, 6576 (1983).
26
J. Nowotny and M. Rekas, Defect Chemistry of (La,Sr)MnO3, J. Am. Ceram.
Soc., 81 [1] 67 80 (1998).
27
J. Fontcuberta, M. A. Crusellas, J. Rodrguez-Carvajal, M. Vallet, J. Alonso, and
J. Gonzalez-Calbet, Mossbauer Study of Vacancy Distribution in CaMn1xFexO3y
(x 0.5, 0.6), J. Solid State Chem., 83, 150 75 (1989).
28
P. D. Battle, C. M. Davison, T. C. Gibb, and J. F. Vente, Structural Chemistry
of SrMn1xFexO3-, x 0.3, J. Mater. Chem., 6 [7] 118790 (1996).

Vous aimerez peut-être aussi